Metric Spaces
Metric Spaces
Metric Spaces
Acharya
Dr. V. V. Acharya
Head, Department of Mathematics,
Fergusson College, Pune-411 004.
Preface
Contents
These are the lecture notes given by me at Fergusson College
n n based on the book by Prof. S. Kumaresan. These notes are useful
1 R, R and C, C 1
for students who want to study metric spaces. An attempt has
1.1 R. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 1
been made to make chapters small so that the student should
1.2 Rn . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 1
understand that concept. There is an emphasis on examples.
1.3 Complex Numbers . . . . . . . . . . . . . . . . . . . . . . . 3
1.4 Cn . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 5
1.5 Some Important Inequalities . . . . . . . . . . . . . . . . . . 6 The aim of the course is to prove the generalization of
the following: theorem: If f : [a, b] R is a continuous
2 Metric Spaces 11 function then f ([a, b]) is a closed and bounded interval.
Suggestions for improvement of notes are welcome. The sugges-
3 Open and closed sets in Metric spaces 15 tions may be sent at vvacharya@gmail.com.
3.1 Open Balls . . . . . . . . . . . . . . . . . . . . . . . . . . . 15
3.2 Open sets . . . . . . . . . . . . . . . . . . . . . . . . . . . . 16
3.3 Closed sets . . . . . . . . . . . . . . . . . . . . . . . . . . . 19
3.4 Sequences in a metric space . . . . . . . . . . . . . . . . . . 20
6 Connected Spaces 47
iii
2 Dr. V. V. Acharya
Let R denote
( the set of real numbers. We define, the absolute value of x by Show that the discriminant of this equation is always non-negative.
x if x 0
|x| = We have seen the following properties of absolute Theorem 1.4 (Cauchy-Schwartz Inequality) Let a1 , . . . , an and
x if x < 0.
b1 , . . . , bn be any real numbers. Then
value function:
(a21 + . . . + a2n )(b21 + . . . + b2n ) (a1 b1 + . . . + an bn )2 (1.1)
Theorem 1.1 1. |x| 0. Further, |x| = 0 if and only if x = 0.
2. |kx| = |k||x| and equality occurs if and only if either
3. |x + y| |x| + |y|. 1. all the ai s are zero or all the bi s are zero or
2. the ai s are proportional to the bi s i.e. there exists k 6= 0 such that
If x, y R then we define distance between x and y by d(x, y) = |x y|.
ai = kbi for all i.
Theorem 1.2
Proof. Follows from exercise 1.1.
1. d(x, y) > 0 if x 6= y and d(x, y) = 0 if x = y. coincide. That is,
d(x, y) 0 for all x, y R and d(x, y) = 0 if and only if x = y. Proof of (3). Note that
n n n n
2. d(x, y) = d(y, x), i.e. distance is symmetric.
kx + yk2
X X X X
= (xi + yi )2 = x2i + yi2 + 2 xi yi
3. If x, y, z are any three points, then d(x, y) d(x, z) + d(z, y). i=1 i=1 i=1 i=1
Xn n
X n
X
Proof. This follows using the properties of the absolute value. x2i + yi2 + 2 |xi yi |
i=1 i=1 i=1
v v
1.2 Rn n
X n
X
u n
X uX
u n
x2i + y 2 + 2t
u
i x2 t y2 i i
n
Note that R = {(x1 , x2 , . . . , xn )|xi R} . Let x = (x1 , x2 , . . . , xn ), y = i=1 i=1 i=1 i=1
(y1 , y2 , . . . , yn ) Rn . We define, x + y = (x1 + y1 , . . . , xn + yn ) and v
u n
v
u n
2
p = (kx1 , . . . , kxn ). Further, the length of x, kxk , is defined by kxk =
kx
uX uX
x2i + t yi2 = (kxk + kyk)2 .
x21 + + x2n . Length of x is also called as norm of x. The norm of x has
t
i=1 i=1
the following properties:
Theorem 1.3 1. kxk 0. Further, kxk = 0 if and only if x = 0. Hence, kx + yk kxk + kyk .
1
Modern Analysis 3 4 Dr. V. V. Acharya
If points A and B on the real axis have co-ordinates a and b respectively, i = (0, 1) is the imaginary unit and i2 = 1. Also x is called the real part
then the distance between them is d(A, B) = |a b|. and y the imaginary part of the complex number z = x + iy and we write
x = Re(z), y = Im(z). Note that since x, y are real numbers, we have
We define the distance between two points in a plane and space as usual:
the distance between the points A(a, b) and P (x, y) in R2 is denoted as x2 0 and y 2 0. (1.5)
d(A, P ) and is defined thus:
p Further, z = x iy is called the conjugate of z and
d(A, P ) = + (x a)2 + (y b)2 . (1.2) p
|z| = + x2 + y 2 (1.6)
Similarly, the distance between the points A(a, b, c) and P (x, y, z) in R3 is
defined as follows: is called the modulus of z, so that z z = |z|2 . By (1.5) we see thatp |z| =
p 0 z =p 0 and also x2 x2 + y 2 and y 2 x2 + y 2 . Hence, |x| x2 + y 2
d(A, P ) = + (x a)2 + (y b)2 + (z c)2 . (1.3) and |y| x2 + y 2 . Thus,
More generally, the distance between the points A(a1 , . . . , an ) and |Re(z)| |z| and |Im(z)| |z| (1.7)
P (x1 , . . . , xn ) in Rn is defined by
p Given any complex numbers z, w we have the triangle inequality
d(A, P ) = + (x1 a1 )2 + + (xn an )2 . (1.4)
|z w| |z| + |w| (1.8)
Note that the distance in Rn as defined above is in fact given by d(A, P ) =
kx ak . Further, it has the following basic properties: which also implies the inequality
Theorem 1.5 1. d(A, P ) > 0 if A, P are distinct points and d(A, P ) = ||z| |w|| |z w| (1.9)
0 if A, P coincide. That is, d(A, P ) 0 always and d(A, P ) = 0 if
and only if A, P coincide. Since, |i| = 1 it follows from (1.8) that
2. d(A, P ) = d(P, A), i.e. distance is symmetric. |x + iy| |x| + |iy| = |x| + |i||y| = |x| + |y|. Hence,
3. If A, B, P are any three points, then d(A, P ) d(A, B) + d(B, P ). |z| |Re(z)| + |Im(z)|. (1.10)
Proof. Note that d(A, P ) = kx ak . Hence, using Theorem 1.3 we get These simple results are of great importance.
the result. Consider rectangular coordinate axes OX, OY in a plane. Then we can
Or If A(a, b), P (x, y) R2 , then property (1.2) says that AP = 0 if and interpret the points in the plane in two ways, depending on our point of
only if AP 2 = 0 if and only if (x a)2 + (y b)2 = 0 if and only if x = a view as follows:
and y = b. Also, if A, P are distinct, then either x 6= a or y 6= b i.e. either To every point in the plane corresponds a unique ordered pair (x, y) of
(x a)2 > 0 or (y b)2 > 0 so that AP > 0. In the similar way, we can co-ordinates of P and also to P there corresponds a unique complex number
prove the general case. Prove it! z = x + iy. In both cases the correspondence is one-to-one. Therefore in
the first case we identify the plane with the Cartesian product R2 = R R
1.3 Complex Numbers and call it the real plane. In the second case we identify the plane with the
set C and call it the complex plane (or Argand plane or Gaussian plane).
Let C denote the set complex numbers. Throughout this section x,y,a,b, Thus the point P (x, y) in the real plane corresponds to the point P (z) in
c,d, will denote real numbers. Recall that complex number z is an ordered the complex plane. Also, in the complex plane, the X and Y axes are called
pair of real numbers, z = (x, y), and is also written as z = x + iy. Here the real axis and the imaginary axis respectively.
Modern Analysis 5 6 Dr. V. V. Acharya
Further, if points A, B represent complex numbers z = x + iy and 2. d(A, P ) = d(P, A), i.e. distance is symmetric.
w = a + ib in the complex plane C, we define the distance d(A, B) between
them to be |z w|, i.e. 3. If A, B, P are any three points, then d(A, P ) d(A, B) + d(B, P ).
Proof. Note that d(A, P ) = kx ak . Hence, using Theorem 1.6 we get
the result.
p
|z w| = d(A, B) = + (x a)2 + (y b)2 . (1.11)
the result for n = m 1. Suppose 0 < a1 a2 . . . am . We note that f (x0 ) is the minimum of f on (0, ). Now, 1
the result is true if a1 = am . Suppose a1 < am . Then clearly 1 p
xp0 yq y p1 yq 1
Thus the result is true for n = m and the induction is complete. We Equality holds if and only if C1 |ak |p = C2 |bk |q for 1 k n for some
also observe that when the ai s are unequal, the arithmetic mean is strictly non-zero constants C1 and C2 .
greater than the geometric mean. Hence, if the arithmetic mean and the |ai | |bi |
Proof. Let x = ,y = . Then using (1.14), we get
geometric mean are equal then all the ai s must be equal. Finally, applying kakp kbkq
this result to the n positive numbers 1/a1 , . . . , 1/an we see that Gn Hn ,
with equality if and only if a1 = . . . = an . |ai | |bi | 1 |ai |p 1 |bi |q
p + (1.18)
kakp kbkp p kakp q kbkqq
Theorem 1.10 (Youngs Inequality) Let x, y be non-negative real num- Summing (1.18) for i = 1 to i = n, we get
1 1
bers. Let p > 1 and q be defined in such a way that + = 1 holds. Then n n
! n
!
p q X |ai | |bi | 1 X |ai |p 1 X |bi |q
p + q .
kakp kbkq p kakp q kbkq
xp yq i=1 i=1 i=1
xy + . (1.14) Simplifying, we get
p q n
X |ai ||bi | 1 1
p q + ,
Equality holds if and only if x = y . kak p kbk p p q
i=1
Proof. Fix y > 0. Define the function
whence we get (1.17). It is easy to see that the equality holds if and only
xp
y q if equality holds in Youngs inequality i.e. if and only if
f (x) = + xy for x > 0.
p q |ai |p |bi |q
p = q for 1 k n.
kakp kbkq
The derivative of f (x) is f (x) = xp1 y. Equating f (x) to 0, we get the
1 11 1 1 p1 p 1
critical point as x0 = y p1 . Now,f (x0 ) = (p 1)xp2 > 0, as p > 1. Thus, p
+
q
=1 q
= p
q= p1
=1+ p1
.
Modern Analysis 9 10 Dr. V. V. Acharya
Thus, equality holds if and only if C1 |ak |p = C2 |bk |q for 1 k n for some
non-zero constants C1 and C2 . n
X
|ai ||ai + bi |p1 kakp
|ai + bi |p1
q (using 1.17)
Now,
Remark 1.2 If we take p = 2 then q = 2. In this case, we get i=1
n
!1/q
p1 q
X X
|ai ||bi | kak2 kbk2 , for all a, b Kn . = kakp |ai + bi |
i i=1
n
!1/q
n
X
Thus, we get Cauchy Schwartz inequlity in K . = kakp |ai + bi |(p1)q (1.21)
i=1
!1/q
With the notation as in Holders Inequality, we have the following: n
X
= kakp |ai + bi |p
Theorem 1.12 (Minkowskis Inequality) Let 1 p . Then i=1
= kakp ka + bkp/q
p (1.22)
n
ka + bkp kakp + kbkp for a, b K . (1.19) n
p/q
X
Similarly, |bi ||ai + bi |p1 kbkp ka + bkp (1.23)
Equality holds if and only if there exist some non-zero constants C1 and C2 i=1
such that C1 a = C2 b.
Proof. We will consider three cases. Using (1.20),(1.22) and (1.23), we get
Case 1. p = . For a, b Kn , we have ka + bk = max{|ai + bi | : p p/q p/q
i = 1 . . . n}. Now, using triangle inequality, we get |ai + bi | |ai | + |bi | ka + bkp kakp ka + bkp + kbkp ka + bkp . (1.24)
kak + kbk . Thus,
Dividing both sides of (1.24) by ka + bkp/q
p , we get (1.19.)
2
ka + bk kak + kbk . Remark 1.3 1. Recall that, in an inner product space, we used Cauchy
Pn Scwartz inequality to establish triangle inequality for norm.
Case 2. p = 1. For a, b Kn , we have ka + bk1 = i=1 |ai + bi |. Now,
using triangle inequality, we get |ai + bi | |ai | + |bi |. Thus, 2. Minkowskis Inequality for p = 2 gives us triangle inequalty in the
inner product space Kn .
ka + bk1 kak1 + kbk1 .
Case 3. 1 < p < . It is easy to see that the result holds if either a or b
equals 0. Hence, we may assume that neither a nor b equals zero. Now,
n n
p
X X
ka + bkp = |ai + bi |p = |ai + bi ||ai + bi |p1
i=1 i=1
n
X n
X
|ai ||ai + bi |p1 + |bi ||ai + bi |p1 (1.20) 1
i=1 i=1
2 We use the following facts: 1. p = q(p1) and q
= p1
p
and 2. p pq = p(1 q1 ) = 1.
12 Dr. V. V. Acharya
d(x, y) > 0 (x, y X, x 6= y) Exercise 2.1 Are the following objects metric spaces, or not?
d(x, y) = d(y, x) (x, y X) 1. The Euclidean plane with points x = (xl , x2 ), but with distance de-
fined by d(x, y) = |x1 y1 |.
d(x, y) d(x, z) + d(y, z) (x, y, z X) (Triangle inequality)
If d is a metric for X, then the ordered pair (X, d) is called a metric space. 2. The set of cities in the United States that have airports, with
d(A, B) = scheduled air travel time from A to B.
Example 2.1 1. Let M = R. Define d(x, y) = |x y|. It is easy to see
that (R, d) is a metric space. 3. The positive real numbers with d(x, y) = x/y.
11
Modern Analysis 13 14 Dr. V. V. Acharya
5. Let C be the set of all complex numbers. For x, y C define 16. Let (X, d) be a metric space. Show that the function d : X X R
2|x y| defined by d (x, y) = inf{1, d(x, y)}, x, y X is also a metric.
d(x, y) = p p . Show that (C, d) is a metric
{1 + |x|2 } {1 + |y|2 }
space. 17. For a metric space (X, d), prove that |d(x, y) d(x, z)| d(y, z).
Interpret this relation with usual metric in the plane.
6. (X, d) is a metric space. Define d(x, y) = C d(x, y) for all x, y X
and 0 < C < 1. Show that d is also a metric on X. 18. If M denotes the set of all n n matrices over the reals, examine
whether the following real valued function d defined on M M is a
d(x, y)
7. If d(x, y) is a metric on X, show that (x, y) = is also a metric on M : d(A, B) = | det A det B|, A, B M.
1 + d(x, y)
metric on X. 19. Define : R R R by (x, y) = |x+ y|. Is a metric for R? Justify!
2 2
8. Let the mapping d : R R R be defined by 20. Let (X, d1 ), (Y, d2 ) be metric spaces and M = X Y. Define
p
d(x, y) = (x1 x2 )2 + (y1 y2 )2 ((x1 , y1 ), (x2 , y2 )) = max{d1 (x1 , x2 ), d2 (y1 , y2 )}
2 2 2 ((x1 , y1 ), (x2 , y2 )) = d1 (x1 , x2 ) + d2 (y1 , y2 )
where x = (x1 , y1 ) R , y = (x2 , y2 ) R . Prove that (R , d) is a
metric space.
q
and d((x1 , y1 ), (x2 , y2 )) = d21 (x1 , x2 ) + d22 (y1 , y2 )
9. Let the mapping d : R2 R2 R be defined by d(x, y) = |x1
x2 | + |y1 y2 | where x = (x1 , y1 ) R2 , y = (x2 , y2 ) R2 . Prove that Show that , as well as d are metric on M.
(R2 , d) is a metric space. We call as a product metric.
11. Let d1 and d2 be two metrics for a non-empty set X. Then d defined
by d(x, y) = d1 (x, y) + d2 (x, y) is also a metric for X.
12. Let (X, d) be a metric space. Show that for arbitrary x, y, z, t in X,
|d(x, y) d(z, t)| d(x, z) + d(y, t).
13. Show that the set Rn of all n-tuples of real numbers with d(x, y) =
n
X
[ (x1 y1 )2 ]1/2 x, y Rn is a metric space.
1
14. Show that the set of continuous real valued functions defined on [0, 1]
with d(f, g) = sup |f (x) g(x)| is a metric space.
0x1
b b b b b b
O a O a a a+
Figure 3.1 Figure 3.2
2
Figure 3.1 shows the open ball B((a, b); r) in (R , d2 ) while Figure 3.2 shows
the open ball B((a, b); d ) in (R2 , d ).
Exercise 3.1Consider
the metric
space
[0, 1]
with usual
metric.
1 1 1 1 3 1
Determine B ; 1 , B 0; ,B ; ,B ; B(0; 2) in C
2 2 4 2 4 2
15
Modern Analysis 17 18 Dr. V. V. Acharya
Proof. Let x B(a; r) and s = d(a, x). Suppose t is a positive real number Proof. If x G, then there exist an open interval B containing x such
less that r s. Consider B(x; t). If y B(x; t) then that B G. 1 Let
Corollary 3.1 The intersection of finite number of open sets is open. Solution 2. Let Gn = B( 12 , 12 + n1 ), n N. Then each Gn is an open set
in R as each Gn is an open ball and Gn = [0, 1] which is not an open set
Theorem 3.6 Every open set G R can be written as G = In where in R.
I1 , I2 , . . . are a finite number or a countable number of open intervals which 1 Note that an open ball in R is an open interval.
are mutually disjoint. (That is Im In = if m 6= n.) 2 We do not discard the possibility that Ix may be an unbounded interval.
Modern Analysis 19 20 Dr. V. V. Acharya
3.3 Closed sets Definition 3.6 Let E be a subset of the metric space M. A point x M
is called a limit point of E if there is a sequence {xn } of points of E which
Definition 3.4 Let F be a subset of the metric space M. We say that F converges to x. The set E of all limit points of E is called the closure of E.
is a closed subset of M if F c is an open set.
Remark 3.3 Every point x of E is a limit point of E. For the sequence
Theorem 3.7 If F1 and F2 are closed subsets of M then F1 F2 is also x, x, . . . converges to x. Thus, if x E then x E, that is E E.
closed.
Remark 3.4 Note that every cluster point of E is a limit point of E.
Proof. Note that (F1 F2 )c = F1c F2c . Now F1c F2c is an open set as
However, the converse does not hold.
F1c , F2c are open and finite intersection of open sets is an open set.
Example 3.5 Let E = (0, 1). Note that 0 and 1 are limits points of E.
Using induction, we get However, 0
/ E and 1
/ E. Hence, E is a proper subset of E.
Corollary 3.2 Union of finitely many closed sets is closed.
3.4 Sequences in a metric space
Note that arbitrary union of closed sets need not be closed.
1 Definition 3.7 Let (M, d) be a metric space. A function s with domain
Example 3.3 Consider Fn = . Since, Fn is singleton and singletons N and having values in M is called an infinite sequence in M. The value
n
[ of s at n, namely s(n), is called the nth term of the sequence and is often
are closed each Fn is a closed set. However, E = Fn is not a closed set denoted as sn . Also, then the sequence is denoted as {sn }nN or simply as
n=1 {sn }. The set of all terms of a sequence, namely {sn |n N}, is called the
as 0 is a limit point of E. range of the sequence {sn }. If all the terms sn are real numbers, then {sn }
is called a real sequence.
T 3.8 If F be any family of closed subsets of a metric space M,
Theorem
then F is also closed. Definition 3.8 Let (M, d) be a metric space and {sn }
n=1 be a sequence
F F of points in M. We say that sn approaches L M as n approaches infinity
F )c =
S c
F . Since, F c is an open set and the
T
Proof. Note that ( if > 0 there is a positive integer r such that
F F F F
arbitrary union of open sets is an open set, the result follows.
d(sn , L) < , for every n r. (3.1)
Example 3.4 Give an example of a metric space (X, d) such that every In this case we write
proper subset of X is both open and closed.
Solution. In discrete metric space (X, d), singleton sets are open as if lim sn = l or lim sn = l or n L as n
n
x X then B(x, 12 ) = {x}. Since arbitrary union of open sets is open,
and we say that {sn }
n=1 is convergent in M to the point L.
every set is open. Thus, if E X then E is open. But E c is open as well.
Thus, E is closed as well as open. Thus, every proper subset of X is both Theorem 3.9 If the sequence {sn }
n=1 is convergent, then its limit is
open and closed. unique.
Definition 3.5 Let E be a subset of the metric space M. A point a M Proof of this theorem is left as an exercise.
is called a cluster point of E if for every h > 0 there exists a point x of E Definition 3.9 Let (M, d) be a metric space and {sn }
n=1 be a sequence
such that 0 < d(a, x) < h. of points in M. We say that sn is a Cauchy sequence if > 0 there is a
positive integer r such that
Remark 3.2 In other words, a M is said to be a cluster point of E if
every open ball B(a; h) contains a point of E distinct from a. d(sn , sm ) < , for all m, n r. (3.2)
Modern Analysis 21 22 Dr. V. V. Acharya
Theorem 3.10 Let (M, d) be a metric space and {sn }n=1 be a sequence Proof. Since, E E we need to prove that E E. Let x E. To show
of points in M. Then {sn }
n=1 is a Cauchy sequence. that x E, it suffices to show that any open ball B[x; r] contains a point
of E. x E, the ball B[x; r] contains a point y E. Let d(x, y) = s.
The converse of this theorem is not true. There are metric spaces in which
Choose a positive number t such that t < r s. Since, y E, the ball
Cauchy sequences need not be convergent. For example, let M = (0, 1)
1 B[y; t] contains a point z E. Now,
with usual metric and sn = . It is easy to see that {sn } is a Cauchy
n+1 d(x, z) d(x, y) + d(y, z) < s + t < s + (r s) = r.
sequence but it is not convergent.
Hence, z B[x; r]. Thus, every open ball B[x; r] contains a point of E.
Definition 3.10 Let (M, d) be a metric space. We say that (M, d) is a
Hence, x is a limit point of E, that is, x E.
Complete metric space if every Cauchy sequence in M is convergent in M.
Theorem 3.13 In any metric space (M, d) the sets M and are both
Example 3.6 If {sn } is a Cauchy sequence in (R, d) then {sn } is eventually closed.
constant and hence convergent. That is, there exist a positive integer M Proof. Since the set M contains all its limit points and that has no limit
such that for all n M, xn = xn+1 = . . . . points (and hence contains all its limit points), both M and are closed.
1
If {sn } is a Cauchy sequence in (R, d) then for = there exist a
2 Theorem 3.14 Let G be an open subset of the metric space M. Then
1 G = M G is closed. Conversely, if F is a closed subset of M, then
positive integer M such that d(sn , sm ) < for all n, m M. Hence,
2 F = M F is open.
d(sn , sm ) = 0 for all n, m M, that is, for all n M, sn = sn+1 = . . . .
Proof. Let G be an open subset of the metric space M. If x G, then
In fact, if (M, d) is a discrete metric space, then every Cauchy sequence
there exists r > 0 such that B[x; r] G. Hence, B[x; r] contains no point
is constant. Thus, every discrete metric space is a complete metric space.
of G . Hence, x cannot be a limit point of G . Thus, no point of G is a limit
Theorem 3.11 Let E be a subset of the metric space M. Then the point point of G , so G contains all its limit points and is thus closed.
x M is a limit point of E if and only if every open ball B[x; r] about x Now, suppose F is a closed subset of M. If y F then as F is closed y
contains at least one point of E. is not a limit point of F. Hence, there exist r > 0 such that B[y; r] contains
Proof. Suppose x is a limit point of E. Then there is a sequence {xn } n=1
no point of F. That is, B[y; r] F . Hence, F is open.
of points of E that converges to x. If B[x; r] is any open ball about x, then
Example 3.8 Note that Z is a closed subset of (R, d) as Z =
S
(n, n+ 1)
there exists M N such that for all n M, d(xn , x) < . Hence, B[x; r] nZ
contains a point of E. and (n, n + 1) is an open set and hence Z is open. Hence, Z is a closed set.
Conversely, let x M and suppose every open ball B[x; r] about x
Solution 2. An alternate way to do this, is to show that Z has no cluster
contains at least one point of E. Then for every n N, the open ball
1 point. If x R Z is a cluster point then we write x = [x] + , 0 < < 1.
B[x; 1/n] contains a point xn E. Since, d(xn , x) < for all n N, the Let r = 12 min{, 1 }. Now, B(x, r) contains no point of Z. Hence, x is
n
sequence {xn }n=1 converges to x. Hence, x is a limit point of E. not a cluster point of Z. Hence Z is a closed set. 3
Example 3.7 Let (M, d) be a metric space and x M. Then {x} is a Definition 3.11 Let M be a metric space. The subset A of M is said to
closed subset of M. For the only sequence of points of {x} is x, x, x, . . . , be dense in M if A = M.
and hence x itself is the only limit point of {x}. Thus, {x} contains all its
Remark 3.5 1. We note that R {0} is an open set of R which is also
limit points and is therefore closed.
dense. In fact, we can remove finitely many points from R and the
Theorem 3.12 If E is any subset of a metric space M, then E is closed. resulting set is dense. It is easy to see that R Z is dense in R.
That is E = E. 3 Equivalently, one may show that every Cauchy sequence in Z is convergent.
Modern Analysis 23 24 Dr. V. V. Acharya
2. Note that as every irrational number is limit of sequence of rationals, Exercise 3.6 1. Show that in a metric space, the complement of a sin-
Q is dense in R. (usual metric) gleton set is open.
3. We note that Qc = R Q (also) is dense in R. 2. Prove that in a metric space, the complement of a finite set is open.
4. A is dense in M if every point in M is a limit point of A. 3. Find the closure of the following subsets of R with usual metric
(a) A = {1/x|x N} (b) Set of integers (c) Q
Example 3.9 Note that the discrete metric space (R, d) has no dense sub- (d) (0, 1) (e) (0, 1] (f) [0, 1).
set except R itself as every subset of (R, d) is both open and closed. Thus, 4. Give an example of a metric space in which
if A R then A = A. Thus, if A 6= R then A cannot be dense in (R, d). (a) every set is both open and closed
(b) to show that intersection of an infinite number of open sets need
Example 3.10 Let S = {n + m 2|a, b Z}. Let a, b R such that a < b. not be open
Then there exists an s S such that a < s < b. Thus, S is dense in R.
(c) of an infinite class of closed sets whose union is not closed
Solution.
If x, y S and k Z then x y, kx S.4 If n + m 2 = (d) of a set which is neither closed nor open
n + m 2 then n = n and m = m . 5 For, if m 6= m then 2 = mnn m , a (e) to show that the intersection of an infinite number of neighbour-
contradiction. hood of a set, need not be a neighbourhood of that set.
Let Sm = m 2 [m 2]. Thus, 0 Sm < 1.
Claim 1. If m 6= m then Sm = Sm
. For if, Sm = Sm then Solutions
m 2 [m 2] = m 2 [m 2] then [m 2] + m 2 = [m 2] + m 2. 1. Let x0 M. If y 6= x0 then r = d(x0 , y) > 0. Note that x0
/ B(y, r/2).
Hence, {x0 }c is open.
Thus, m = m .
2. Follows from 1 and from the fact that finite intersection of open sets
Hence, we conclude that {Sm |m Z} is an infinite subset of S [0, 1).
is open.
Given > 0, we partition [0, 1) into k sub-intervals so that each sub-
interval has length < . At least one of these sub-intervals must contain 3. (a) A {0} (b) Z (c) R (d), (e) and (f) [0, 1].
two distinct elements, say Sm and Sm . Without loss of generality, we may
assume that Sm < Sm . Thus, 0 < Sm Sm < .
Claim 2. Let > 0 such that < b a. Then there exists an n Z such
that a < n < b.
We may assume that 0 < a < b. By Archimedean property there exists
N such that a < N . Using well ordering principle, choose n to be the least
positive integer k such that k > a. Thus, (n 1) < a n. We claim that
n < b. For otherwise, b a n (n 1) = , a contradiction.
ba
We take = . Then there exists s S such that 0 < s < . Hence
2
there exists an integer n such that a < ns < b. 6 Since, ns S, the claim
is proved.
4 Note that S is an integral domain. It is also true
5 This
that S is a Euclidean domain.
means that S is a free Z-module and {1, 2} is a basis.
6 Take s = as in claim 2.
26 Dr. V. V. Acharya
2. lim (f (x) g(x)) = L N. Example 4.1 Note that f (x) = x2 is a continuous function from R onto
xa
[0, ) while g(x) = sin(x) is a continuous function from R onto [1, 1].
3. lim f (x)g(x) = LN.
xa 1
Theorem 4.5 Every function from (R, d) into a metric space is contin-
f (x) L uous on (R, d).
4. lim = , provided N 6= 0.
xa g(x) N 1 discrete metric
25
Modern Analysis 27 28 Dr. V. V. Acharya
(
Proof. If {xn }n=1 is a sequence of points in (R, d) converging to a, then x if x is rational,
{xn } is a Cauchy sequence. Hence, using Example 3.6, there exist M N Example 4.3 Let f : R R defined by f (x) =
1 x if x is irrational.
such that xn = xn+1 = xn+2 = . . . for all n M, i.e., {xn } is a constant 1
sequence after the M th term. Hence, {f (xn )} is a constant sequence after We will show that f is discontinuous at every point a R except a = .
2
the M th term and hence convergent. Hence, f is continuous. Let a be an irrational number and {xn } be a sequences of rational numbers
( which converges to a. Note that f (a) = 1 a and f (xn ) = xn and hence
1 if x is rational, {f (xn )} will converge to a 6= f (a). Hence, f is discontinuous at every
Example 4.2 Let : R R defined by (x) =
0 if x is irrational. irrational number a R.
We will show that is discontinuous at every point a R. Let {xn } and Let b be a rational number and {xn } be a sequences of irrational num-
{yn } be sequences of rational and irrational numbers respectively which bers which converges to b. Note that f (b) = b and f (xn ) = 1 xn and
converge to a. Note that {(xn )} will converge to 1 and {(yn )} will con- hence {f (xn )} will converge to 1 b. Now, b = 1 b if and only if b = 12 .
verge to 0. Hence, is discontinuous at every a R. Hence, f is discontinuous at every rational number except 12 .
Now we discuss the continuity of f at 12 . Note that
4.2 Equivalent Definitions of Continuity
f (x) f ( 1 ) = x 1
2 2
Theorem 4.6 Let (X, d1 ) and (Y, d2 ) be metric spaces and f be a function
from X to Y and x X.Then the following are equivalent: Hence, given > 0, take = . Now, |f (x)f ( 21 )| < whenever |x 12 | < .
1
1. f is continuous at x. Hence, f is continuous only at .
2
2. Given > 0 there exists > 0 such that d2 (f (x), f (a)) < whenever Example 4.4 2
Let f : R+ R defined by
d(x, a) < .
0 if x is irrational,
3. Given an open set V containing f (x) Y, we can find an open set U
1
containing x such that f (U ) V. f (x) = if x is a rational number of the form m
n , where
n
m, n > 0 are integers and gcd(m, n) = 1.
Proof. Suppose (2) holds i.e. given > 0 there exist > 0 such that
d2 (f (x), f (a)) < whenever d1 (x, a) < . If {xn }
n=1 is a sequence of points Claim: f is continuous at every irrational number and discontinuous at
in X converging to a, then corresponding to > 0 there exist M N such every rational number.
that d1 (xn , a) < for all n M. Hence, d2 (f (xn ), f (a)) < whenever Let a > 0 be a rational number and {xn } be a sequence of irrational
n M. Hence, the sequence {f (xn )} n=1 of points in M2 converges to f (a). numbers which converges to a. Now {f (xn )} will converge to 0 while f (a) >
Conversely, assume that (1) holds. Suppose (2) does not hold. Then 0. Hence, f is discontinuous at a. Thus, f is discontinuous at every rational
there exists > 0 such that for every > 0 there exists x M1 such that number.
1
d1 (x, a) < but d2 (f (x), f (a)) . In particular, for = there exists On the other hand, if b is an irrational number and > 0 then by the
n 1
xn M1 such that d1 (xn , a) < but d2 (f (xn ), f (a)) . Thus, xn a as Archimedean property, there is a natural number n0 such that < .
n0
n but f (xn ) does not tend to f (a), a contradiction. Hence, (2) holds. There are only a finite number of rationals with denominator less than
(1) if and only if (3) is left as an exercise. n0 in the interval (b 1, b + 1). Hence, > 0 can be chosen so small
that the neighbourhood (b , b + ) contains no rational numbers with
Also, one can show that (1) (2) (3) (1). We leave this as an
exercise. 2 This function was introduced by K.J.Thomae in 1875.
Modern Analysis 29 30 Dr. V. V. Acharya
denominator less than n0 . It then follows that for |x b| < , x R+ , 2. The set G M1 is open if and only if its image f (G) M2 is open.
1
we have |f (x) f (b)| = |f (x)| < . Hence f is continuous at every 3. The set F M1 is closed if and only if its image f (G) M2 is closed.
n0
irrational number.
Thus, f is continuous at every irrational number and discontinuous at Proof. Follows using Theorem 4.7 and Theorem 4.8.
every rational number. Hence, f is Riemann integrable over [0, 1].
Example 4.5 Let X, Y be metric spaces and f, g : X Y be continuous
Exercise Set functions. Let E = {x|f (x) 6= g(x)}. Show that E is an open set.
Solution. Let x E. Hence, f (x) 6= g(x). Thus, = d(f (x), g(x)) > 0.
1. Give an example of a function which is continuous on R and whose Let B1 = B(f (x); /2) and B2 = B(g(x); /2). Thus, B1 B2 = . Since,
range is (0, ) f and g are continuous functions, there exist 1 > 0 and 2 > 0 such that
2. Show that every function from Z into a metric space is continuous. f (B(x; 1 )) B1 and g(B(x; 2 )) B2 . Thus, if = min{1 , 2 }, then
f (B(x; )) B1 and g(B(x; )) B2 . Since, B1 B2 = we get that
3. Let (X, d) be a metric space. If is the product metric defined on B(x; ) E. Hence, E is an open set.
X X by ((x1 , y1 ), (x2 , y2 )) = max{d1 (x1 , x2 ), d2 (y1 , y2 )} show that
d is a continuous function from X X to R. Or
Theorem 4.7 Let (M1 , d1 ) and (M2 , d2 ) be metric spaces and f : M1 Let x be a limit point of E c and (xn ) be a sequence of points in E c con-
M2 be a function. Then f is continuous on M1 if and only if f 1 (G) is an verging to x. Since, f and g are continuous functions, f (xn ) and g(xn )
open subset of M1 whenever G is an open subset of M2 . converges to f (x) and g(x). Since xn E c , f (xn ) = g(xn ) for every n.
Proof. Suppose f is continuous on M1 . We wish to show that if G Thus, f (x) = g(x) and E c is a closed set. Hence, E is an open set.
is open in M2 then f 1 (G) is an open subset of M1 . If x f 1 (G)
then y = f (x) G. Since G is an open set there exist s > 0 such Example 4.6 Show that one can construct a continuous function from a
that B[y; s] G. Since, f is continuous at x, corresponding to this s, metric space (M1 , d1 ) to a metric space (M2 , d2 ) which maps closed sets to
there exists r > 0 such that d1 (x1 , x) < r implies d2 (f (x1 ), f (x)) < s, closed sets but does not map open sets to open sets.
that is, f (B[x; r]) B[y; s] or equivalently, f 1 (B[y; s]) B[x; r]. Thus, Solution. This is easy. For example, consider a constant function from R
B[x; r] f 1 (B[y; s]) f 1 (G). Hence, f 1 (G) is an open set. to R.
Conversely, let a M1 and b = f (a). Let s > 0 and B = B[a; s].
Example 4.7 Show that there exist a continuous function from a metric
Now B is an open set. Hence, f 1 (B) is an open set in M1 . Further,
space (M1 , d1 ) to a metric space (M2 , d2 ) which maps open sets to open
a f 1 (B). Since, f 1 (B) is an open set in M1 there exists r > 0 such
sets but does not map closed sets to closed sets.
that B[a; r] f 1 (B). Thus, f is continuous. 1
Solution. Consider f : R+ R where f (x) = . Note that f is one-one,
Theorem 4.8 Let (M1 , d1 ) and (M2 , d2 ) be metric spaces and f : M1 x
continuous function. It is easy to see that f maps open sets to open sets.
M2 be a function. Then f is continuous on M1 if and only if f 1 (F ) is a However f (N) is not a closed set in R as 0 is a cluster point even though N
closed subset of M1 whenever F is a closed subset of M2 . is a closed set in R+ .
Proof. Follows from Theorem 4.7 and Theorem 3.14
Definition 4.4 Let (M1 , d1 ) and (M2 , d2 ) be metric spaces and f : M1
Theorem 4.9 Let (M1 , d1 ) and (M2 , d2 ) be metric spaces and f : M1
M2 be a one-one and onto function. If both f and f 1 are continuous (on
M2 be a one-one and onto function. Then if f has one of the following
M1 and M2 , respectively) then we call f a homeomorphism from M1 onto
properties then it has them all.
M2 . If a homeomorphism from M1 onto M2 exists, then we say that M1
1. Both f and f 1 are continuous (on M1 and M2 , respectively). and M2 are homeomorphic.
Modern Analysis 31 32 Dr. V. V. Acharya
Remark 4.2 Theorem (4.9) tells us that if f : M1 M2 be a one-one and is a one-one function. Further f g = IdJ and g f = IdI . Hence, f and
onto function such that G is open in M1 if and only if its image f (G) M2 g are onto functions and are inverses of each other. Hence, I and J are
is open then f is a homeomorphism. Equivalently, if f : M1 M2 be a homeomorphic.
one-one and onto function such that F is closed in M1 if and only if its
Example 4.13
image f (F ) M2 is closed then f is a homeomorphism.
Example 4.8 Note that the metric spaces [0, 1] and [0, 2] with absolute
value metric are homeomorphic. For if f (x) = 2x, then f is a homeomor- Consider the function f : (1, 1) R
phism from [0, 1] onto [0, 2]. x
defined by f (x) = . Note that
1 x2
f (x) is a rational function i.e. it is
)
Example 4.9 Note that the metric spaces (0, 1) and (0, with absolute
a ratio of two polynomials. Hence, it
value metric are homeomorphic. For if f (x) = tan x , then f is a
2 is continuous at all points where de-
homeomorphism from (0, 1) onto (0, ). nominator does not vanish. In this
case, the denominator never vanishes as
Example 4.10 Note that the metric spaces R and (0, ) with absolute x (1, 1). Further, it is easy to see
value metric are homeomorphic. For if f (x) = ex , then f is a homeomor- that f (x) is one-one and onto function.
phism from R onto (0, ). Can you determine the inverse of f ? Is
f 1 continuous?
Example 4.11 Note that the metric spaces (0, ] and R with absolute
value metric are homeomorphic. For if g(x) = log x, then g is a homeomor- x
phism from (0, ) onto R. Graph of y =
1 x2
Exercise 4.1 1. Show that (0, 1) and R are homeomorphic.
Example 4.12 Show that any two closed and bounded intervals in R are
homeomorphic. 2. Show that R+ and R are homeomorphic.
Let I = [a, b] and J = [c, d]. There are two possibilities:
Case I: a = c. Thus, I = [a, b] and J = [a, d]. Define f : I J by 3. Show that (0, 1] and [1, ) are homeomorphic.
f (x) = a + (d a)/(b a)(x a). Note that f is a polynomial. Hence, 4. Show that R2 and C are homeomorphic.
f is continuous. Further, f (x) = f (y) a + (d a)/(b a)(x a) =
a + (d a)/(b a)(y a) x = y. Thus, f is a one-one function. Define Solutions
g : J I by g(x) = a + (b a)/(d a)(x a). Note that g is a polynomial. 1. Define f : (0, 1) R by f (x) = x 12 . f maps (0, 1) to ( 12 , 12 ).
Hence, g is continuous. Further, g(x) = g(y) a + (b a)/(d a)(x a) = Define g : ( 12 , 12 ) R by g(y) = tan(y). Note that f and g are
a + (b a)/(d a)(y a) x = y. Thus, g is a one-one function. Further homeomophisms. By taking their composition, we get the function.
f g = IdJ and g f = IdI . Hence, f and g are onto functions and are
inverses of each other. Hence, I and J are homeomorphic. 2. Note that f (x) = log(x) is a one-one onto continuous function from
R+ to R and the inverse map is x 7 ex .
Case II: a 6= c. Define f : I J by f (x) = a + (d b)/(c a)(x a). Note
that f is a polynomial. Hence, f is continuous. Further, f (x) = f (y) 1
3. Note that f (x) = is a one-one onto continuous function from (0, 1]
a + (d b)/(c a)(x a) = a + (d b)/(c a)(y a) x = y. Thus, f is a x
one-one function. Define g : J I by g(x) = c+ (b d)/(a c)(x c). Note to [1, ) and the inverse map is f 1 (x) = x1 .
that g is a polynomial. Hence, g is continuous. Further, g(x) = g(y) 4. Define f : R2 C by f (x, y) = (x, y). This is a homeomorphism.
c + (b d)/(a c)(x c) = c + (b d)/(a c)(y c) x = y. Thus, g
34 Dr. V. V. Acharya
Example 5.1 Note that every subset of a bounded set is bounded. Also,
every subset of (R, d) is a bounded set as d(x, y) 1 for all x, y R.
Chapter 5 Definition 5.5 Let (M, d) be a metric space. We say that the subset A
of M is totally bounded if, given > 0, there exists a finite number of
subsets A1 , . . . , An of M such that diam Ak < , 1 k n and such that
Compact Metric Spaces A
[n
Ak .
k=1
5.1 Compact Set Example 5.2 If A is a totally bounded subset of the metric space (M, d)
then every subset of A is totally bounded.
Definition 5.1 Let M be any set, the[family F of subsets of A of M is Example 5.3 Show that every finite subset of a metric space (M, d) is
said to form a covering of M if M A.
totally bounded.
AF
Let A = {a1 , . . . , an }. Then Ai = B[ai , /3] cover A and diam Ai < .
Definition 5.2 Let (M, d) be a metric space and A be a subset of M. Hence, A is totally bounded.
The family
[ F of {Ui |i I} of M is said to form an open covering of A if Theorem 5.1 If the subset A of metric space (M, d) is totally bounded,
A Ui and each Ui F is an open set. then A is bounded.
iI
[ Proof. If A is totally bounded, then there exist nonempty subsets of M, say
n
If J I such that A Ui , we then say that {Ui |i J} is a subcover [
iJ
A1 , . . . , An , such that diam Ak < 1, 1 k n and such that A Ak .
of the given open cover of A. k=1
S
If there exists a finite subset J I such that A iJ Ui , then we say For every k, 1 k n, let ak Ak and D = d(a1 , a2 ) + d(a2 , a3 ) + +
that the given cover admits a finite subcover. d(an1 , an ). Let x, y A. Since, Ai s cover A there exist Ai and Aj such
that x Ai and y Aj for some i and j. We may assume that i j. Now,
Definition 5.3 Let (M, d) be a metric space and A be a subset of M. A d(x, y) d(x, ai ) + [d(ai , ai+1 ) + + d(aj1 , ai )] + d(aj , y)
is said to be compact if every open cover has a finite subcover.
d(x, ai ) + D + d(aj , y).
Since, diam Ai < 1 and diam Aj < 1, d(x, ai ) < 1 and d(aj , y) < 1. Hence,
5.2 Bounded and Totally Bounded sets
d(x, y) < 1 + D + 1 = D + 2, for all x, y A.
Definition 5.4 Let (M, d) be a metric space. We say that the subset A of
M is bounded if there exists a positive number L such that Hence, A is a bounded set.
Theorem 5.2 Every bounded subset of (R, d) is totally bounded.
d(x, y) L x, y A. Proof. Let A be a bounded subset of R. Then A [L, L] for some L > 0.
Given > 0, by Archimedean property, there exist an integer n such that
If A is bounded we define the diameter of A, denoted diam(A), as
n( ) > 2L and A is covered by
2
diam A = sup d(x, y).
x,yA [L, L + ], [L + , L + 2 ], . . . , [L + (n 1) , L + n ]
2 2 2 2 2
If A is not bounded we write diam(A) = . Hence, A is a totally bounded set.
33
Modern Analysis 35 36 Dr. V. V. Acharya
n
Example 5.4 If we take M = [0, 1] with usual metric, then we note that [
that A Ak . For every k, 1 k n, let ak Ak . Then {a1 , . . . , an }
every subset of M is totally bounded.
k=1
is -dense in A. Hence, if A is totally bounded then A has a finite -dense
Theorem 5.3 Every bounded subset of (Rn , d) is totally bounded.
subset.
Proof. We will write the proof for n = 2. Let A be a bounded subset
Conversely, if {x1 , . . . , xn } is 3 -dense in A, then {B(x1 , 3 ), . . . , B(xn , 3 )}
of R2 . Then A [L, L] [L, L] for some L > 0. Given > 0, by
forms a covering of A by sets of diameter < . Hence, A is totally bounded.
Archimedean property, there exist an integer n such that n( ) > 2L. We
3
divide [L, L] [L, L] into small squares of length /3 by dividing [L, L] Theorem 5.6 Let (M, d) be a metric space. The subset A of the metric
on x-axis as well as y-axis as space (M, d) is totally bounded if and only if every sequence of points of A
contains a Cauchy subsequence.
[L, L + ], [L + , L + 2 ], . . . , [L + (n 1) , L + n ] Proof. Suppose A is totally bounded and (xn ) n=1 be a sequence of points
3 3 3 3 3 of A. We want to show that (xn ) has a Cauchy subsequence. Since the
n=1
Thus, we get n2 squares whose diagonal (and hence diameter) is of length set A is totally bounded, A can be covered by a finite number of subsets of A
< . Hence, A is totally bounded. of diameter < 1. One of these sets, call it A1 , must contain xn for infinitely
In a similar way, we can write the proof in general case.1 many values of n. Choose an integer n1 such that xn1 A1 . Since, A1 A
and A is totally bounded, it follows that A1 is totally bounded and hence
Theorem 5.4 Every totally bounded subset of the discrete metric space 1
A1 can be covered by a finite number of subsets of A1 of diameter < .
(R, d) contains only finitely many elements. 2
One of the sets, say A2 must contain xn for infinitely many values of n.
Proof. If A is totally bounded subset of (R, d), then there exist nonempty
Choose an integer n2 > n1 such that xn2 A2 . Since, A2 A1 we also
subsets A1 , . . . , An of M such that diam Ak < 1, 1 k n and such that
[n have xn2 A1 . Continuing in this way we obtain, for every natural number
A Ak . Since, diam Ak < 1, Ak contains exactly one element. Hence, n, a subset An of An1 such that diam Ak < 1/k, and xnk Ak . Since,
k=1 xnk , xnk+1 , . . . all lie in Ak , and diam Ak < 1/k, it follows that (xnk ) is a
A has only finitely many elements. Cauchy subsequence of (xn ).
Conversely, suppose every sequence (xn ) n=1 of points of some subset
Definition 5.6 Let A be a subset of the metric space (M, d) and > 0. A of M contains a Cauchy sequence. We wish to show that A is totally
The subset B of A is said to be -dense in A if for every x A there exists bounded. Suppose the contrary. Then there exists > 0 such that A
y B such that d(x, y) < . contains no finite -dense subset. Thus, if x1 A then the set {x1 } is
not -dense in A, so there exist x2 A such that d(x1 , x2 ) . But
Thus, B is -dense in A if each point of A is within distance from some
the set {x1 , x2 } is not -dense in A. Hence, there exist x3 A such
point of B.
that d(x1 , x3 ) , d(x2 , x3 ) . Continuing in this way we can con-
Theorem 5.5 Let (M, d) be a metric space. The subset A of M is totally struct a sequence {xn } n=1 of points of A such that d(xj , xk ) for any
bounded if and only if for every > 0 there exists a finite subset {x1 , . . . , xn } j, k N, j 6= k. But then (xn ) n=1 has no Cauchy subsequence, which con-
We now show that M is complete. Let (xn ) be a Cauchy sequence in We claim that lim xn = l. Let > 0 be given. Now there exist M N
n
M. Then for every k N there exists nk such that d(xn , xnk ) < k1 for all such that
n > nk . Let |xn xM | < /2 for n M
1
Uk = {x M |d(x, xnk ) > }.
k using (5.2).
Note that Uk is complement of B[xnk , k1 ]. Hence, Uk is open. Now xn / Uk
Example 5.5 Note that (R, d), where d is a discrete metric is complete
for n > nS k . Hence, no finite subcover of Uk s cover X. For, if they did,
metric spaces. In fact, every discrete metric space is a complete metric
say M = m i=1 Ui , we take n > max{n1 , . . . , nm }. Then xn
/ Uk for any k
space as every Cauchy sequence in a discrete metric space is eventually
with 1 kS m. This implies that {Uk } cannot cover M. Thus there exists
constant.
x M i=1 Ui . But then d(x, xnk ) < k1 . Hence, xnk x. Since, (xn ) is
a Cauchy sequence in M, we see that (xn ) also converges to x. Thus, X is 1. Are the following subspaces of R with the usual metric complete?
complete. Give reasons for your answer.
(1) The set of all positive reals. (2) The set of rationals.
5.3 Complete Metric spaces (3) [1, 2] {3, 4}. (4) The set of integers. (5) The set of irrationals.
Definition 5.7 Let (M, d) be a metric space. M is said to be a complete Example 5.6 Show that (0, 1) with usual metric is not a complete metric
metric space if every Cauchy sequence of points in M converges to a point space.
in M . 1
Let an = . We show that (an ) is a Cauchy sequence. Let > 0.
n
Theorem 5.8 R is a complete metric space. 1 1
Choose M N such that < . If m, n > M then d(am , an ) = | m n1 | <
Proof. Let (xn ) be a Cauchy sequence in R. Let us take > 0. By the M 2
1 1
definition of Cauchy sequence, there exits N () such that m + n < . Hence, (an ) is a Cauchy sequence. This sequence converges to
0 but 0
/ (0, 1). Thus, (0, 1) is not a complete metric space.
|xn xm | < for m N (), n N (). (5.1)
Theorem 5.9 If (M, d) is a complete metric space and A is a closed subset
In particular, of M then (A, d) is a complete metric space.
|xn xN () | < for n N (). (5.2) Proof. Let {xn } n=1 be a Cauchy sequence of points in (A, d). Since, A
M, {xn }n=1 is a Cauchy sequence of points in (M, d). But (M, d) is a
If we take = 1 then there exist N = N (1) such that
complete metric space. Hence, {xn } n=1 converges to some x M. But x is
|xn xN | < 1 for n N. (5.3) a limit point of A because x is the limit of a sequence of points in A. Since,
A is closed, x A. Hence, every Cauchy sequence in A converges to some
This implies that xn (xN 1, xN + 1) for all n N. Let point in A. Hence, (A, d) is a complete metric space.
S = {x R| there exist infinitely many n such that xn x}. Theorem 5.10 Let (M, d) be a complete metric space. For each n N,
let Fn be a closed bounded subset of M such that
Note that xN 1 S and hence S is non-empty. Further, xN + 1 is an
upper bound of S. If this is not true, then there exists y S such that F1 F2 . . . Fn Fn+1 . . . , (5.4)
y > xN + 1 and there are infinitely many n such that xn y. This implies and diam Fn 0 as n . (5.5)
that xn > xN + 1 for infinitely many n, a contradiction. Thus, S is a
non-empty subset of R and xN + 1 is an upper bound of S. Hence, by LUB
T
axiom, S has a least upper bound, say l. Then Fn contains precisely one point.
n=1
Modern Analysis 39 40 Dr. V. V. Acharya
T
Proof. For each n N, let an Fn . Then by (5.4), By (5.10), there is precisely one point x An . Now, since F is a
covering of M, there is a set G F such that x G. But G is open as F
an , an+1 , . . . all lie in Fn . (5.6) is an open covering and so there exists r > 0 such that B(x, r) G. Since,
r > 0 there exists a natural number N such that N1 < r.
Given > 0 there exists by (5.5), an integer M N such that Since, x AN and diam (AN ) < N1 < r, we get AN B(x, r) G.
diam (FM ) < . Hence, G alone covers AN , a contradiction. This proves the theorem.
Theorem 5.12 The metric space (M, d) is compact if and only if every
Now, aM , aM+1 , . . . all lie in FM . For n, m M, we have
sequence of points in M has a subsequence converging to a point in M.
d(an , am ) diam (FM ) < . Proof. Let (M, d) be compact metric space and that {xn } n=1 be a sequence
of points in M. Since, M is totally bounded, the sequence {xn } n=1 has a
This proves that {an } is a Cauchy sequence in M. Since, (M, d) is a complete Cauchy subsequence {xnk } k=1 . But {xnk
}
k=1 converges to a point in M
metric space, {an } is convergent to some a M. By (5.6), a is a limit point since M is a complete metric space. Thus, if M is compact then every
T sequence of points in M has a subsequence converging to a point in M.
of the closed set Fn for each n. Hence, a Fn for each n. Hence, a Fn . Conversely assume that every sequence in M has a convergent subse-
n=1
If b M such that a 6= b then d(a, b) > diam Fn for n sufficiently large. quence. Then, M is totally bounded. To show that M is complete we must
T
T show that every Cauchy sequence {xn } n=1 in M converges to a point in M.
Hence, b cannot be in Fn . Hence, Fn contains precisely one point.
n=1 n=1 By assumption, {xn } n=1 has a subsequence {xnk }
n=1 which converges to a
point x M. Since, {xn }n=1 is a Cauchy sequence, given > 0 there exists
M N such that d(xm , xn ) < /2 for all m, n M. Since, {xnk } n=1 which
5.4 Equivalent conditions for Compact metric spaces
converges to a point x M there exists r N such that d(xnk , x) < /2
Theorem 5.11 If (M, d) is totally bounded and complete metric space, for all k r, (i.e. nk nr ). If p = max{nr , M } then np p nr and for
then M is compact. all k p, we have
Proof. Suppose the contrary. Then for some open covering F , no finite
d(xk , x) d(xk , xnp ) + d(xnp , x) < /2 + /2 = .
number of sets in F form a covering of M. Now M is totally bounded and
hence may be written as the union of a finite number of bounded subsets Hence, the Cauchy sequence {xn }n=1 in M converges to a point x M.
each of whose diameter is less than 1. But one of these subsets, say A1 , Hence, M is a complete metric space. Since, M is also totally bounded, M
cannot be covered by a finite number of sets in F . (Otherwise all of M is a compact metric space.
would be covered by a finite number of sets in F .) We now consider A1 .
Note that diam A1 = diam A1 and A1 is a closed subset of M. Further, Corollary 5.1 If A is a closed subset of the compact metric space (M, d)
diam A1 < 1 and A1 cannot be covered by a finite number of sets in F . then the metric space (A, d) is also compact.
Since A1 is itself totally bounded, the same reasoning shows the existence Proof. Let {xn }
n=1 be a sequence of points in A. Since, A M, {xn }n=1 is
of a subset A2 whose diameter is less than 1/2 and which cannot be covered a sequence of points in M. Since, (M, d) is a compact metric space, {xn }
n=1
by a finite number of sets in F . Thus, A2 A1 , diam A2 < 1/2 and A2 has a subsequence converging to a point x M. This implies that x is a
cannot be covered by a finite number of sets in F . Continuing in this way, limit point of A. Since, A is closed, x A. Thus, every sequence in A has
for every n N, we can show the existence of An M such that a subsequence converging to a point in A. Hence, the metric space (A, d) is
also compact.
A1 A2 A2 , diam An < 1/n
Theorem 5.13 Let A be a subset of the metric space (M, d). If (A, d) is
and such that An cannot be covered by a finite number of sets in F . compact then A is a closed subset of the metric space (M, d).
Modern Analysis 41 42 Dr. V. V. Acharya
Proof. Let x be a limit point of A. Then there is a sequence sequence Conversely, assume that whenever F is\ a family of closed subsets of M
{xn }
n=1 of points of A converging to x. But then {xn }n=1 is a Cauchy with the finite intersection property, then F 6= . We want to show that
sequence in A, and so, since A is complete, {xn }n=1 converges to a point in F F
A. This point must be x (as the limit of a sequence is unique, if exists) and M is compact. Consider G to be an open cover of M. If G does not have
so x A. Thus A contain all its limit points and so A is closed.2 a finite subcover of M , then this means that whenever G1 , G2 , . . . , Gn are
elements of G, M 6 G1 G2 Gn . Hence, there exists x M such that
Definition 5.8 A family F of subsets of a set M is said to have the finite x / G1 G2 Gn , that is, x (G1 G2 Gn )c . If Fi denote (Gi )c
intersection property if the intersection of any finite number of sets in F is then Fi is a closed set as each Gi is an open set. If F denote the set of all
never empty. Gc where G G then F is a family of closed sets S having finite intersection
c
( GG G)c ,
S
property. However, as G is an open cover, M GG G, M
Thus, a family F of subsets of a set M is said to have the finite intersection i.e. = M c ( GG Gc ). As Gc = F, we get F F F = , a contradiction.
T T
property if whenever F1 , . . . , Fn F , then F1 F2 Fn 6= . Hence, G has a finite subcover of M. That is, M is compact.
Example 5.7 Note that the family of all closed intervals [1/n, 1/n] has
the finite intersection property. Also, the family of open intervals (0, 1/n) 5.5 Continuous functions on compact metric space
has the finite intersection property.
Theorem 5.15 Let f be a continuous function from the compact metric
Theorem 5.14 The metric space M is compact if and only if whenever F space M1 into the metric space M2 . Then the range f (M1 ) of f is also
is\a family of closed subsets of M with the finite intersection property, then compact.
F 6= . Proof 1. Let G be any open covering of f (M1 ). Since, f is continuous
F F and G is an open set, f 1 (G) is an open subset of M1 . The family of all
Proof. Suppose first that M is a compact metric space and that F is a such sets f 1 (G) for G G is therefore an open covering of M1 . Since M1
family of closed subsets of M with the finite intersection property. We want is compact,a finite number of these sets, say f 1 (G1 ), . . . , f 1 (Gn ), form
to show that \ a covering of M1 . But then G1 , . . . , Gn form a covering of f (M1 ), and so
F 6= . (5.7) f (M1 ) is compact.
F F Proof 2. Let {yn } n=1 be any sequence in f (M1 ). For every n N let xn
For each F F let G = F = M f, and let G be the family of all such M1 such that f (xn ) = yn . Since, M is compact {xn } n=1 has a subsequence
open sets G. xnk which converges to a point, say x of M. Since f is continuous on
If F1 , . . . , Fn F , then using de Morgans laws, we get M1 f (xnk ) converges to f (x), that is, f (yk ) converges to f (x). Thus, any
sequence {yn } n=1 in f (M1 ) has a convergent subsequence and hence f (M1 )
F1 F2 Fn = M (G1 G2 Gn ). (5.8) is compact.
Corollary 5.3 Let f be a real valued function continuous on a closed and Proof. We want to show that f 1 is continuous. Let F be a closed subset
bounded interval in R. Then f must be bounded. of M1 . We want to show that the inverse image of F under f 1 is closed
Proof. Since, every closed and bounded interval is compact, we get the in M2 . But the inverse image of F under f 1 is f (F ). Since, M1 is
required result. compact and F is closed, F itself is compact. Since, f is continuous f (F ) is
compact and hence a closed subset of M2 = f (M1 ). Hence, f is continuous.
Theorem 5.16 Let f be a continuous function from the compact metric
space M into R. Then f attains a maximum value at some point of M. Example 5.9 Let g : [0, ) [0, ) defined by g(x) = x. We want to
Also, f attains a minimum value at some point of M. prove that g is a continuous function. Let N N. Then f : [0, N ] [0, N 2 ]
Proof 1. Note that f is a bounded function. Let = sup f (x) and = defined by f (x) = x2 is a continuous function from the compact metric
xM space to compact metric space. Hence, f is a homeomorphism. Hence, f 1
inf f (x). Hence, and are limit points of f (M ). Since, f (M ) is compact, is a continuous function on [0, N 2 ]. But f 1 is the restriction of g to [0, N 2 ].
xM
f (M ) is a closed subset of R. Hence, f (M ) and f (M ). Hence, there Hence, g is continuous.
exist a, b in M such that f (a) = and f (b) = .
Proof 2. With the notation as in Proof 1, suppose / f (M ), then the 5.7 Uniform Continuity
function g(x) = f (x) is a continuous function on M and g never takes
the value 0 on M. Hence, 1/g is continuous on M. Since, M is compact, Definition 5.10 Let (M1 , d1 ) and (M2 , d2 ) be metric spaces and f be a
1 1 function from M1 to M2 . We say that f is uniformly continuous on M1
1/g is bounded (being continuous). Thus, = N for some
g(x) L f (x) if given > 0 there exists > 0 such that d2 (f (x), f (a)) < whenever
1 1 d1 (x, a) < .
N 0. and all x M. Hence, f (x) L for all x M. Hence, L
N N
is an upper bound for f (M ), a contradiction. Hence, f (M ). Exercise 5.1
Similarly, we can prove that f (M ).
1. Show that if (M, d) is a discrete metric space then every subset is
Corollary 5.4 Let f be a real valued function continuous on a closed and both open and closed. Further, show that its every compact subset is
bounded interval in [a, b]. Then f attains a maximum and a minimum value finite.
at points of [a, b].
Proof. Since, every closed and bounded interval is compact, we get the 2. Show that every finite subset is compact.
required result. 3. Show that {B(x, 1/2)}xX is an open cover of X.
Example 5.8 Let C[0, 1] denote the collection of all real valued bounded 4. Show that {B(x, n)}nN is an open cover of X.
continuous functions defined on [0, 1]. Define the norm of f C[0, 1] by,
kf k = sup{|f (x)| : x [0, 1]} Then prove that the mapping d : C[0, 1] R, 5. Show that a closed subset of a compact metric space is compact.
defined by d(f, g) = kf gk = sup{|f (x) g(x)| : x [0, 1]} is a metric for
6. Show that a closed subset of a complete metric space is complete.
C[0, 1].
7. If (xn ) is a Cauchy sequence and it has a convergent subsequence
converging to x then show that (xn ) converges to x.
5.6 Continuity of the inverse function
8. Show that following sets are closed subset of R3 . Check whether each
Theorem 5.17 If f is a one-one continuous function from the compact of them is compact?
metric space M1 onto the metric space M2 , then f 1 is continuous on M1 ,
and hence f is a homeomorphism. (a) {(x, y, z)|x2 = y 2 + z 2 }
Modern Analysis 45 46 Dr. V. V. Acharya
47
Modern Analysis 49 50 Dr. V. V. Acharya
Theorem 6.6 (Connected subsets of R) A set J R is connected if assume that an = 1. If a0 = 0 then p(0) = 0 i.e 0 is a root of p(x). We may
and only if J is an interval. assume that a0 6= 0. Define
Proof. If J R is not an interval, then there exist x, y J and a z between
p(x) an1 a0
( that z
x and y such / J. Consider the function f : R {z} {1} defined q(x) = i.e. q(x) = 1 + + + n.
1 if s < z xn x x
by f (s) = . Then f is a non-constant continuous function.
1 if s > z. For x > 1, we have
Hence, f |J is a non-constant continuous function onto {1}. Hence, J R a
n1 a0 |an1 | |a0 |
is not connected. |q(x) 1| = + + n + + n
x x |x| |x |
To prove the converse, let J be any non-empty interval. Let x, y J.
n
Since J is an interval, [x, y] J. Further, [x, y] is a connected set. Thus, J |an1 | + + |a0 | A X
= , where A = |ai |
has the property that two of its points lie in a connected set. Hence, J is |x| |x| i=0
a connected set by the above lemma.
Thus, if |x| > max{1, 2A}, then |q(x) 1| < 12 . Hence, q(x) > 0 if |x| >
Theorem 6.7 (Intermediate Value Theorem) Let f : [a, b] R be a max{1, 2A}. If we now choose > max{1, 2A}, then q() and q() are
continuous function. Assume that y is a point between f (a) and f (b). Then both positive. But then p() is positive and p() is negative. Hence by
there exists x [a, b] such that f (x) = y. intermediate value theorem there exists [, ] such that p() = 0.
Proof. Note that [a, b] is connected. Since f is a continuous function on
[a, b], its image under f, that is, f [a, b] is connected. Since, any connected Example 6.7 Show that the annulus A = {(x, y)|1 < x2 + y 2 < 4} is
subset of R is an interval and f [a, b] is connected, f [a, b] is an interval, say connected.
J. Now, f (a), f (b) J. By the definition of an interval, the point y J, Solution 1. Note that [1, 2) and [0, 2] are connected subsets of R. Hence,
that is y f [a, b]. Therefore, there exists x [a, b] such that f (x) = y. [1, 2) [0, 2] are connected subsets of R2 . Define f : [1, 2) [0, 2] R2 by
f (r, ) = (r cos , r sin ). Note that f is a continuous function and image of
Theorem 6.8 (Existence of nth roots) Let [0, ) and n N. f is A. Since, [1, 2) [0, 2] are connected subsets of R2 , A is a connected
Then there exists a unique x [0, ) such that xn = . subset of R2 .
Proof. If = 0 then x = 0 and 0n = 0. So we assume that > 0. Solution 2. If we pick up any two points in A, then we can join them by
Consider the continuous function f : [0, ) R defined by f (t) = tn . a polygonal path which lies entirely in A. Hence, A is connected. 2
By Archimedean property of |R there exits a natural number N such that
N > . We have Theorem 6.10 Let X and Y be connected metric spaces. Then the prod-
uct space X Y is connected.
0 = f (0) < and f (N ) = N n N > . Proof. Let f : X Y {1} be a continuous function. Let (x0 , y0 )
X Y be fixed. Let (x, y) X Y be an arbitrary point in X Y. If we
Hence, using Intermediate value theorem to the interval [0, N ], we see that
show that f (x, y) = f (x0 , y0 ) then we are done.
there exists x [0, N ] such that xn = .
To prove the above claim, let us first observe that iy : X X Y
Suppose there exists y > 0 such that y n = . We may assume that
defined by iy (x) = (x, y) is a continuous function. Similarly, the map
x < y. Hence, xn < y n < , a contradiction. Hence, x is unique.
ix : Y X Y defined by ix (y) = (x, y) is a continuous function. Therefore
for every point y Y, the subset X {y} = {(x, y)|x X} is a connected
Theorem 6.9 Any polynomial with real coefficients and of odd degree has
subset of X Y ; similarly, the subset {x} Y = {(x, y)|x X} is a
a real root.
connected subset of X Y for every point x X.
Proof. Let p(x) = an xn + an1 xn1 + + a1 x + a0 , where ai R and
an 6= 0 and n is odd. Since the roots of p(x) and a1n p(x) are same, we may 2 This solution is better than the first, as we can easily generalize this.
Modern Analysis 53 54 Dr. V. V. Acharya
Now the point (x, y0 ) lies in both X {y0 } and {x} Y. The restriction
of f to either of these sets are continuous and hence constants as the subsets
X {y} and X {y0 } are connected. Hence, f (x, y0 ) = f (x, y) for every
y Y and f (x, y0 ) = f (x0 , y0 ) for every x X. Thus, f (x, y) = f (x, y0 ) =
f (x0 , y0 ) for every x X and y Y. Thus, f is constant. Hence, X Y is
connected.
It follows that
|v| kyk |y|,
thus proving the other inequality, and completing the proof of our assertion.
55 56